Вы находитесь на странице: 1из 38

Fluid mechanics

Chapter 1

1-1 Introduction
Fluid mechanics is the study of fluids either in motion (fluid
dynamics) or at rest (fluid statics) and the subsequent effects of the fluid
upon the boundaries, which may be either solid surfaces or interfaces
with other fluids. Both gases and liquids are classified as fluids, and the
number of fluids engineering applications is enormous: breathing, blood
flow, swimming, pumps, fans, turbines, airplanes, ships, rivers,
windmills, pipes, missiles, icebergs, engines, filters, jets, and sprinklers,
to name a few. When you think about it, almost everything on this planet
either is a fluid or moves within or near a fluid.

Fig. 1.1 Flow processes occur in many ways in our natural environment

In fluid mechanics, however, one is not content with the formulation


of the laws by which fluid movements are described, but makes an effort
beyond that to find solutions for flow problems, i.e. for given initial and
boundary conditions. To this end, three methods are used in fluid
mechanics to solve flow problems:
(a) Analytical solution methods (analytical fluid mechanics):
Analytical methods of applied mathematics are used in this field to solve
the basic flow equations, taking into account the boundary conditions
describing the actual flow problem.
(b) Numerical solution methods (numerical fluid mechanics):
Numerical methods of applied mathematics are employed for fluid flow
simulations on computers to yield solutions of the basic equations of fluid
mechanics.
(c) Experimental solution methods (experimental fluid mechanics):
This sub-domain of fluid mechanics uses similarity laws for the
transferability of fluid mechanics knowledge from model flow
investigations. The knowledge gained in model flows by measurements is
transferred by means of the constancy of known characteristic quantities
of a flow field to the flow field of actual interest.
From the point of view of fluid mechanics, all matter consists of only
two states, fluid and solid. The difference between the two is perfectly
obvious to the layperson, and it is an interesting exercise to ask a
layperson to put this difference into words. The technical distinction lies
with the reaction of the two to an applied shear or tangential stress. A
solid can resist a shear stress by a static deformation; a fluid cannot.
Any shear stress applied to a fluid, no matter how small, will result in
motion of that fluid. The fluid moves and deforms continuously as long
as the shear stress is applied. As a corollary, we can say that a fluid at rest
must be in a state of zero shear stress, a state often called the hydrostatic
stress condition in structural analysis. In this condition, Mohr’s circle for
stress reduces to a point, and there is no shear stress on any plane cut
through the element under stress.
Given the definition of a fluid above, every layperson also knows that
there are two classes of fluids, liquids and gases. Again the distinction is
a technical one concerning the effect of cohesive forces. A liquid, being
composed of relatively close-packed molecules with strong cohesive
forces, tends to retain its volume and will form a free surface in a
gravitational field if unconfined from above. Free-surface flows are
dominated by gravitational effects. Since gas molecules are widely
spaced with negligible cohesive forces, a gas is free to expand until it
encounters confining walls. A gas has no definite volume, and when left
to itself without confinement, a gas forms an atmosphere which is
essentially hydrostatic. Gases cannot form a free surface, and thus gas
flows are rarely concerned with gravitational effects other than buoyancy.
Figure 1.2 illustrates a solid block resting on a rigid plane and
stressed by its own weight. The solid sags into a static deflection, shown
as a highly exaggerated dashed line, resisting shear without flow. A free-
body diagram of element A on the side of the block shows that there is
shear in the block along a plane cut at an angle 𝜃𝜃 through A. Since the
block sides are unsupported, element A has zero stress on the left and
right sides and compression stress 𝜎𝜎 = -p on the top and bottom. The
liquid and gas at rest in Fig. 1.1 require the supporting walls in order to
eliminate shear stress. The walls exert a compression stress of -p. The
liquid retains its volume and forms a free surface in the container. If the
walls are removed, shear develops in the liquid and a big splash results. If
the container is tilted, shear again develops, waves form, and the free
surface seeks a horizontal configuration, pouring out over the lip if
necessary. Element A in the gas is also hydrostatic and exerts a
compression stress -p on the walls.
Fig. 1.2 A solid at rest can resist shear. (a) Static deflection of the solid; (b) equilibrium
and Mohr’s circle for solid element A. A fluid cannot resist shear. (c) Containing walls
are needed; (d) equilibrium and Mohr’s circle for fluid element A.

1-2 Physical Basics:


1-2-1 Solids & Fluids
All substances of our natural and technical environment can be
subdivided into solid, liquid and gaseous media, on the basis of their state
of aggregation. This subdivision is accepted in many fields of engineering
in order to reveal important differences concerning the properties of the
substances. This subdivision could also be applied to fluid mechanics,
but, it would not be particularly advantageous. It is rather recommended
to employ fluid mechanics aspects to achieve a subdivision of media, i.e.
a subdivision appropriate for the treatment of fluid flow processes. To
this end, the term fluid is introduced for designating all those substances
that cannot be classified clearly as solids. Hence, from the point of view
of fluid mechanics, all media can be subdivided into solids and fluids, the
difference between the two groups being that solids possess elasticity as
an important property, whereas fluids have viscosity as a characteristic
property. Shear forces imposed on a solid from outside lead to inner
elastic shear stresses which prevent irreversible changes of the positions
of molecules of the solid. When, in contrast, external shear forces are
imposed on fluids, they react with the build-up of velocity gradients,
where the build-up of the gradient results via a molecule-dependent
momentum transport, i.e. momentum transport through fluid viscosity.
Thus elasticity (solids) and viscosity (liquids) are the properties of matter
that are employed in fluid mechanics for subdividing media. However,
there are a few exceptions to this subdivision, such as in the case of some
of the materials in rheology exhibiting mixed properties. They are
therefore referred to as visco-elastic media. Some of them behave such
that for small deformations they behave like solids and for large
deformations they behave like liquids.
For further subdivision of fluids, it is recommended to make use of their
response to normal stresses (or pressure) acting on fluid elements. When
a fluid element reacts to pressure changes by adjusting its volume and
consequently its density, the fluid is called compressible. When no
volume or density changes occur with pressure or temperature, the fluid is
regarded as incompressible although, strictly, incompressible fluids do
not exist. However, such a subdivision is reasonable and moreover useful
and this will be shown in the following derivations of the basic fluid
mechanics equations. Indeed, this subdivision mainly distinguishes
liquids from gases.

1-2-2 Continuum Mechanics


As all matter consists of molecules or aggregations of molecules, all
macroscopic properties of matter can be described by molecular
properties. Hence it is possible to derive all properties of fluids that are of
importance for considerations in fluid mechanics, from properties of
molecules, i.e. macroscopic properties of fluids can be described by
molecular properties. However, a molecular description of the state of
matter requires much effort owing to the necessary extensive formalism
and moreover the treatment of macroscopic properties would remain
unclear. A molecular-theoretical treatment of fluid properties would
hardly be appropriate to supply application-oriented fluid mechanics
information in an easily comprehensible form. For this reason, it is more
advantageous to introduce quantities of continuum mechanics for
describing fluid properties. The connection between continuum
mechanics quantities, introduced in fluid mechanics, and the molecular
properties should be considered, however, as the most important links
between the two different ways of description and presentation of fluid
properties. Some properties of the thermodynamic state of a fluid, such
as density ρ, pressure P and temperature T, are essential for the
description of fluid mechanics processes and these can easily be
expressed in terms of molecular quantities.
However, there are some important domains in fluid mechanics where
continuum considerations are not appropriate, e.g. the investigation of
flows in highly diluted gas systems. No clear continuum mechanics
quantities can be defined there for the density and pressure with which
fluid mechanics processes can be resolved. The required spatial resolution
of the fluid mechanics considerations does not provide, due to the
dilution, sufficient numbers of molecules for the necessary establishment
of the mean values of the considered continuum properties. Hence there
are insufficient molecules available in the considered δV for the
introduction of the continuum mechanics quantities. When treating such
fluid flows, priority has to be given to the molecular theory rather than
continuum mechanics considerations. In the present introduction to fluid
mechanics, the domain of flows of highly diluted gases is not dealt with,
so that all required considerations can take place in the terminology of
continuum mechanics. For continuum mechanics considerations,
molecular effects, e.g. within the conservation laws for mass, momentum
and energy, are presented in integral form, i.e. the molecular structure of
the considered fluids is not neglected but taken into consideration in the
form of integral quantities.

1-2-3 Dimensions and Units


Since in our study of fluid mechanics we will be dealing with a variety
of fluid characteristics, it is necessary to develop a system for describing
these characteristics both qualitatively and quantitatively. The qualitative
aspect serves to identify the nature, or type, of the characteristics (such as
length, time, stress, and velocity), whereas the quantitative aspect
provides a numerical measure of the characteristics. The quantitative
description requires both a number and a standard by which various
quantities can be compared. A standard for length might be a meter or
foot, for time an hour or second, and for mass a slug or kilogram. Such
standards are called units, and several systems of units are in common use
as described in the following section. The qualitative description is
conveniently given in terms of certain primary quantities, such as length,
L, time, T, mass, M, and temperature 𝜃𝜃.

Table 1.1 Primary Dimensions in SI and BG Systems

Table 1.2 Secondary Dimensions in Fluid Mechanics


1-2-4 Pressure ( P )
The normal force pushing against a plane area, divided by the area, is
the average pressure intensity. The pressure intensity at a point is the
limit of the ratio of normal force to area as the area approaches zero size
at the point. The pressure at a point in the fluid can be expressed
mathematically as follows:
Δ𝐹𝐹
𝑃𝑃 = lim 1.1
Δ𝐴𝐴→0 Δ𝐴𝐴

where Δ𝐹𝐹 is the incremental normal force exerted on the incremental area
Δ𝐴𝐴 by the surrounding fluid particles.
From the point of view of molecular theory, the pressure effect is defined
as the momentum change per unit time felt per unit area, i.e. the force
which the molecules experience and exert on a wall when colliding in an
elastic way with the wall in the considered area. The following relation
holds
1 ���2 = 1 𝜌𝜌𝑢𝑢
���2
𝑃𝑃 = 𝑚𝑚𝑚𝑚𝑢𝑢 1.2
3 3
where m is the molecular mass, n the number of the molecules per unit
volume and mean u the thermal velocity of the molecules.

Fig. 1.3 Fluids exert a compressive force on any surface they contact. (A) A fluid
exerting a force normal to a solid surface. (B) A fluid exerting a compressive force on an
imaginary interior surface. Note that the force exerted on the fluid above and to the
right of the imaginary surface is equal and opposite to the force exerted by the fluid
below and to the left of that same surface.
For the surface forces to be in equilibrium, the sum of the vertical and
horizontal components must be equal to zero. If the forces per unit area,
the pressures on the surface, are denoted by p1, p2, and p3, then the
forces can be written as products of the pressures and the areas, on which
they act. The following sketch shows the prismatic element with the
surface forces indicated. If another geometric shape of the volume
element would have been chosen, the equilibrium condition would always
require the vanishing of the sum of the vertical and horizontal
components of the surface forces.
𝑎𝑎𝑎𝑎𝑎𝑎
𝑝𝑝1 𝑎𝑎𝑎𝑎 − 𝑝𝑝3 𝑐𝑐𝑐𝑐 cos(𝑎𝑎, 𝑐𝑐) − 𝜌𝜌𝜌𝜌 =0
2

𝑝𝑝2 𝑏𝑏𝑏𝑏 − 𝑝𝑝3 𝑐𝑐𝑐𝑐 cos(𝑏𝑏, 𝑐𝑐) = 0


𝑎𝑎 = 𝑐𝑐 cos(𝑎𝑎, 𝑐𝑐) , 𝑏𝑏 = 𝑐𝑐 cos⁡
(𝑏𝑏, 𝑐𝑐)
𝑝𝑝1 = 𝑝𝑝2 = 𝑝𝑝3 = 𝑝𝑝 1.3
For c → 0 the volume forces vanish. It
follows for every point in a fluid which
is in equilibrium, that the pressure p
does not depend on the direction of the surface element on which it acts.
The equilibrium condition for a cylinder with infinitesimally small cross-
sectional area A, and with its axis normal to the positive direction of the
gravitational force, yields the following relation
𝑝𝑝1 𝐴𝐴 = 𝑝𝑝2 𝐴𝐴 ⇒ 𝑝𝑝1 = 𝑝𝑝2 = 𝑝𝑝 1.4

1-2-5 Density ( 𝝆𝝆)


The density represents the mass of the fluid contained within a unit
volume. Mathematically, the density at a point is given as
Δ𝑚𝑚
𝜌𝜌 = lim 1.5
Δ𝐴𝐴→0 Δ𝑉𝑉
where Δ𝑚𝑚 is incremental mass of the fluid surrounding the point and Δ𝑉𝑉
is incremental volume surrounding the point.

1-2-6 Specific Weight ( 𝜸𝜸 )


The specific weight of a fluid, designated by the Greek symbol 𝛾𝛾
(gamma), is defined as its weight per unit volume. Thus, specific weight
is related to density through the equation
𝛾𝛾 = 𝜌𝜌𝜌𝜌 1.6
where g is the local acceleration of gravity. Just as density is used to
characterize the mass of a fluid system, the specific weight is used to
characterize the weight of the system.

1-2-7 Specific Volume ( 𝑽𝑽𝑺𝑺𝑺𝑺 )


The specific volume of a fluid is the volume occupied by a unit mass.
It is, therefore, given by
1
𝑉𝑉𝑆𝑆𝑆𝑆 = 1.7
𝜌𝜌

1-2-8 Specific Gravity ( SG )


The specific gravity of a fluid, designated as SG, is defined as the ratio
of the density of the fluid to the density of water at some specified
temperature. Usually the specified temperature is taken as 4 0𝐶𝐶 and
specified pressure is taken as 76 cm.
𝑆𝑆𝑆𝑆
𝜌𝜌
= 1.8
𝜌𝜌𝐻𝐻2 𝑂𝑂 ( 𝛳𝛳=4 0𝐶𝐶, 𝑃𝑃=76 𝑐𝑐𝑐𝑐 )
1-2-9 Viscosity ( 𝝁𝝁 )
The properties of density and specific weight are measures of the
“heaviness” of a fluid. It is clear, however, that these properties are not
sufficient to uniquely characterize how fluids behave since two fluids
(such as water and oil) can have approximately the same value of density
but behave quite differently when flowing. There is apparently some
additional property that is needed to describe the “fluidity” of the fluid.

Fig. 1.4 Fluid element under force.

The shearing force F acts on the area on the top of the element. This area
is given by 𝐴𝐴 = 𝛿𝛿𝛿𝛿 ∙ 𝛿𝛿𝛿𝛿. W can thus calculate the shear stress ( 𝜏𝜏 ) which
is equal to force per unit area
𝐹𝐹
𝜏𝜏 = 1.9
𝐴𝐴
The deformation which this shear stress cause is measured by the size
of angle Φ and is known as shear strain,
𝑥𝑥 Φ
Φ= , 𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟 𝑜𝑜𝑜𝑜 𝑠𝑠ℎ𝑒𝑒𝑒𝑒𝑒𝑒 𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠 (𝑅𝑅) = .
𝑦𝑦 𝑡𝑡

𝑥𝑥 𝑢𝑢
𝑅𝑅 = = , 𝑢𝑢 𝑖𝑖𝑖𝑖 𝑡𝑡ℎ𝑒𝑒 𝑣𝑣𝑣𝑣𝑣𝑣𝑣𝑣𝑣𝑣𝑣𝑣𝑣𝑣𝑣𝑣 𝑜𝑜𝑜𝑜 𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝 𝑎𝑎𝑎𝑎 𝐸𝐸 1.10
𝑡𝑡𝑡𝑡 𝑦𝑦
𝑢𝑢 𝑢𝑢
⟹ 𝜏𝜏 = 𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐 = 𝜇𝜇 1.11
𝑦𝑦 𝑦𝑦
𝑢𝑢
The term is the change in velocity with y, or the velocity gradient, and
𝑦𝑦
𝑑𝑑𝑑𝑑
may be written in the differential form . The constant of proportionality
𝑑𝑑𝑑𝑑

is known as the dynamic viscosity, 𝜇𝜇, of the fluid


𝑑𝑑𝑑𝑑
𝜏𝜏 = 𝜇𝜇 1.11
𝑑𝑑𝑑𝑑
This is known as Newton's law of viscosity.
Fluids may be classified a Newtonian or non-Newtonian. In
Newtonian fluid there is a linear relation between the magnitude of
applied shear stress and the resulting rate of deformation as shown in Fig.
1.5.

Fig. 1.5 Rheological diagram.


In non-Newtonian fluid there is a nonlinear relation between the
magnitude of applied shear stress and the rate of angular deformation. An
ideal plastic has a definite yield stress and a constant linear relation of 𝜏𝜏
𝑑𝑑𝑑𝑑
to .
𝑑𝑑𝑑𝑑
1-2-10 Kinematic Viscosity (𝝊𝝊 )
The ratio of the coefficient of viscosity to the density is called the
kinematic viscosity, a quantity in which no force is involved. That is
𝜇𝜇
𝜈𝜈 = 1.12
𝜌𝜌

1-2-11 Temperature (𝜽𝜽 )


When two bodies of different heat content are brought into contact
while isolated from all other bodies, there is a flow of heat until the
bodies are in thermal equilibrium. Under this condition they are said to
have a property in common; this property is called temperature.

1-2-12 Thermal conductivity ( K )


The thermal conductivity of a substance has the same mathematical
form as the viscosity coefficient but with different variables. That is,
there is a law known as Fourier's heat conduction law which states that
the conductive heat flow per unit area 𝑞𝑞𝑛𝑛 is proportional to the
temperature decrease per unit distance in a direction normal to the area
𝜕𝜕𝜕𝜕
through which the heat is flowing. Then 𝑞𝑞𝑛𝑛 𝛼𝛼 − , the constant of
𝜕𝜕𝜕𝜕

proportionality is this equation is called the thermal conductivity.


𝜕𝜕𝜕𝜕
𝑞𝑞𝑛𝑛 = −𝐾𝐾 1.13
𝜕𝜕𝜕𝜕
1-2-13 Specific Heat ( 𝑪𝑪𝒗𝒗 , 𝑪𝑪𝒑𝒑 )
The specific heat is defined as the amount of heat required to raise the
temperature of a unit mass of medium by one degree. The heat supplied
per unit mass is
𝐻𝐻 = 𝐶𝐶𝑣𝑣 (𝜃𝜃2 − 𝜃𝜃1 ) = 𝐶𝐶𝑝𝑝 (𝜃𝜃2 − 𝜃𝜃1 ) 1.14

1-2-14 Surface Tension (𝝈𝝈 )


At the interface between a liquid and a gas, or between two immiscible
liquids, forces develop in the liquid surface which cause the surface to
behave as if it were a “skin” or “membrane” stretched over the fluid
mass. Although such a skin is not actually present, this conceptual
analogy allows us to explain several commonly observed phenomena. For
example, a steel needle will float on water if placed gently on the surface
because the tension developed in the hypothetical skin supports the
needle. Small droplets of mercury will form into spheres when placed on
a smooth surface because the cohesive forces in the surface tend to hold
all the molecules together in a compact shape. Similarly, discrete water
droplets will form when placed on a newly waxed surface.
These various types of surface phenomena are due to the unbalanced
cohesive forces acting on the liquid molecules at the fluid surface.
Molecules in the interior of the fluid mass are surrounded by molecules
that are attracted to each other equally. However, molecules along the
surface are subjected to a net force toward the interior. The apparent
physical consequence of this unbalanced force along the surface is to
create the hypothetical skin or membrane. A tensile force may be
considered to be acting in the plane of the surface along any line in the
surface. The intensity of the molecular attraction per unit length along
any line in the surface is called the surface tension and is designated by
the Greek symbol 𝜎𝜎 (sigma). For a given liquid the surface tension
depends on temperature as well as the other fluid it is in contact with at
the interface.
𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠ℎ𝑖𝑖𝑖𝑖𝑖𝑖 𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓
𝜎𝜎 = 1.15
𝑢𝑢𝑢𝑢𝑢𝑢𝑢𝑢 𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙ℎ

1-2-15 Capillarity
Among common phenomena associated with surface tension is the rise
(or fall) of a liquid in a capillary tube. If a small open tube is inserted into
water, the water level in the tube will rise above the water level outside
the tube as is illustrated in Fig. 1.6a. In this situation we have a liquid–
gas–solid interface. For the case illustrated there is an attraction
(adhesion) between the wall of the tube and liquid molecules which is
strong enough to overcome the mutual attraction (cohesion) of the
molecules and pull them up the wall. Hence, the liquid is said to wet the
solid surface.

Fig. 1.6 Effect of capillary action in small tubes. (a) Rise of column for a liquid
that wets the tube. (b) Free-body diagram for calculating column height. (c)
Depression of column for a non-wetting liquid

The height, h, is governed by the value of the surface tension, 𝜎𝜎, the tube
radius, R, the specific weight of the liquid, 𝛾𝛾, and the angle of contact, 𝜃𝜃,
between the fluid and tube. From the free-body diagram of Fig. 1.6b we
see that the vertical force due to the surface tension is equal to
2𝜋𝜋𝜋𝜋𝜋𝜋 𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐 and the weight is 𝛾𝛾𝛾𝛾𝑅𝑅2 ℎ and these two forces must balance
for equilibrium. Thus, 𝛾𝛾𝛾𝛾𝑅𝑅 2 ℎ = 2𝜋𝜋𝜋𝜋𝜋𝜋 𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐.
2𝜎𝜎 𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐
⟹ ℎ= 1.16
𝛾𝛾𝛾𝛾
The angle of contact is a function of both the liquid and the surface. For
water in contact with clean glass 𝜃𝜃 ≈ 0𝑜𝑜 .
Chapter 2

2-1 Pressure and equilibrium


2-1-1 Basic Equation for Pressure Field
Although we have answered the question of how the pressure at a
point varies with direction, we are now faced with an equally important
question—how does the pressure in a fluid in which there are no shearing
stresses vary from point to point? To answer this question considers a
small rectangular element of fluid removed from some arbitrary position
within the mass of fluid of interest as illustrated in Fig. There are two
types of forces acting on this element: surface forces due to the pressure,
and a body force equal to the weight of the element. Other possible types
of body forces, such as those due to magnetic fields, will not be
considered in this text.

Surface and body forces acting on small fluid element


If we let the pressure at the center of the element be designated as p, then
the average pressure on the various faces can be expressed in terms of p
and its derivatives as shown in the figure. We are actually using a Taylor
series expansion of the pressure at the element center to approximate the
pressures a short distance away and neglecting higher order terms that
will vanish as we let δx, δy, and δz approach zero. For simplicity the
surface forces in the x direction are not shown. The resultant surface force
in the y direction is

2.1
Or

2.2

Similarly, for the x and z directions the resultant surface forces are

2.3

The resultant surface force acting on the element can be expressed in


vector form as

2.4
Or

2.5

where 𝑖𝑖̂, 𝑗𝑗̂, and 𝑧𝑧̂ are the unit vectors along the coordinate axes shown in
the figure. The vector forms the pressure gradient and can be written as

2.6
Where

2.7
Thus, the resultant surface force per unit volume can be expressed as

2.8

Since the z axis is vertical, the weight of the element is

2.9

where the negative sign indicates that the force due to the weight is
downward (in the negative z direction). Newton’s second law, applied to
the fluid element, can be expressed as

2.10

Where ∑ 𝛿𝛿𝑭𝑭 represents the resultant force acting on the element, a is the
acceleration of the element, and is the element mass, which can be written
as It follows that

2.11
Or

2.12
And therefore

2.13

This equation is the general equation of motion for a fluid in which there
are no shearing stresses.

For a fluid at rest a = 0, therefore

2.14
or in component form

2.15

2-1-2 Conditions of equilibrium of fluids


Taking the most general case suppose a mass of liquid elastic or non-
elastic homogeneous or heterogeneous to be at rest under the action of
given forces. We want to determine the condition of equilibrium and the
pressure at any point. Let x, y, z be the coordinates reefed to rectangular
axis of any point B of the fluid. Let Q be any other point near it, where
BQ is parallel to the axis of x. take 𝑥𝑥 + 𝛿𝛿𝛿𝛿, 𝑦𝑦, 𝑧𝑧 as the coordinates of Q,
about BQ describe a small prism or cylinder terminated by planes
perpendicular to BQ and let A be the area of the section of the cylinder
perpendicular to its axis and let P the pressure in that side and 𝑃𝑃 + 𝛿𝛿𝛿𝛿
the pressure at the other side. Then the thrust at the point B is PA, and

the thrust at point Q is (𝑃𝑃 + 𝛿𝛿𝛿𝛿)A. If 𝜌𝜌 be the mean density of cylinder


BQ, its mass equal 𝜌𝜌𝜌𝜌𝜌𝜌𝜌𝜌 and 𝑋𝑋𝑋𝑋𝑋𝑋𝑋𝑋𝑋𝑋 will represent the force on BQ
parallel to its axis, if 𝑋𝑋𝑋𝑋𝑋𝑋, 𝑌𝑌𝑌𝑌𝑌𝑌, 𝑍𝑍𝑍𝑍𝑍𝑍 be the components of the forces
acting on a particle 𝛿𝛿𝛿𝛿 of fluid at point x, y, z. Hence, for the equilibrium
of BQ,
(𝑃𝑃 + 𝛿𝛿𝛿𝛿)𝐴𝐴 − 𝑃𝑃𝑃𝑃 = 𝑋𝑋𝑋𝑋𝑋𝑋𝑋𝑋𝑋𝑋
⟹ 𝛿𝛿𝛿𝛿 = 𝜌𝜌𝜌𝜌𝜌𝜌𝜌𝜌
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
= 𝜌𝜌𝜌𝜌, = 𝜌𝜌𝜌𝜌, = 𝜌𝜌𝜌𝜌 2.16
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
where X, Y, and Z are the component of acceleration in the direction of x,
y, z respectively. But
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝑧𝑧
𝑑𝑑𝑑𝑑 = 𝜌𝜌(𝑋𝑋 𝑑𝑑𝑑𝑑 + 𝑌𝑌 𝑑𝑑𝑑𝑑 + 𝑍𝑍 𝑑𝑑𝑑𝑑) 2.17
the equation which determines the pressure.
We now consider what condition must be satisfied by a given distribution
of force in order that it may be capable of maintaining a fluid in
equilibrium. The pressure is a function of x, y, and we know that

𝜕𝜕 2 𝑃𝑃 𝜕𝜕 2 𝑃𝑃
⎧ =
⎪ 𝜕𝜕𝜕𝜕𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕𝜕𝜕𝜕𝜕
⎪ 2
𝜕𝜕 𝑃𝑃 𝜕𝜕 2 𝑃𝑃
= 2.18
⎨ 𝜕𝜕𝜕𝜕𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕𝜕𝜕𝜕𝜕
2

⎪ 𝜕𝜕 𝑃𝑃 = 𝜕𝜕 2 𝑃𝑃
⎩𝜕𝜕𝜕𝜕𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕𝜕𝜕𝜕𝜕

𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕 𝜕𝜕𝜕𝜕
⎧ � �= � �
⎪ 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕

𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕 𝜕𝜕𝜕𝜕
� �= � � 2.19
⎨ 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
⎪ 𝜕𝜕
⎪ 𝜕𝜕𝜕𝜕 𝜕𝜕 𝜕𝜕𝜕𝜕
� �= � �
⎩𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕

𝜕𝜕 𝜕𝜕
⎧ (𝜌𝜌𝜌𝜌) = (𝜌𝜌𝜌𝜌)
⎪ 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕

𝜕𝜕 𝜕𝜕
(𝜌𝜌𝜌𝜌) = (𝜌𝜌𝜌𝜌) 2.20
⎨ 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
⎪ 𝜕𝜕 (𝜌𝜌𝜌𝜌) = 𝜕𝜕 (𝜌𝜌𝜌𝜌)

⎩𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
⎧𝑍𝑍 + 𝜌𝜌 = 𝑌𝑌 + 𝜌𝜌
⎪ 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕

𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑋𝑋 + 𝜌𝜌 = 𝑍𝑍 + 𝜌𝜌 2.21
⎨ 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕

⎪𝑌𝑌 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
+ 𝜌𝜌 = 𝑌𝑌 + 𝜌𝜌
⎩ 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕

𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕


⎧𝑍𝑍 − 𝑌𝑌 = 𝜌𝜌 � − �
⎪ 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕

𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑋𝑋 − 𝑍𝑍 = 𝜌𝜌 � − � 2.22
⎨ 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕

⎪𝑌𝑌 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
− 𝑌𝑌 = 𝜌𝜌 � − �
⎩ 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕

Multiplying by X, Y, Z and adding, we get


𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑋𝑋 � − � + 𝑌𝑌 �𝜕𝜕𝜕𝜕 − � + 𝑍𝑍 �𝜕𝜕𝜕𝜕 − �=0 2.23
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕

as a necessary condition of equilibrium.

2-1-2 Homogeneous Liquids


If the fluid be homogeneous and incompressible, it follows from
equation 2.15 that P must be a perfect differential, in other words the
forces can be represented by the space variation of a potential function.
𝑑𝑑𝑑𝑑 = −𝜌𝜌𝜌𝜌𝜌𝜌
⟹ 𝑃𝑃⁄𝜌𝜌 + 𝑉𝑉 = 𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐 2.24
And this is the condition of equilibrium for a homogeneous fluid.
2-1-3 Fluid at Rest (under the action of
gravity)
Taking the axis of z vertical, and measuring z downwards X = 0, Y = 0
and Z = g. In this case of homogeneous fluid we obtain
𝑑𝑑𝑑𝑑 = 𝑔𝑔𝑔𝑔 𝑑𝑑𝑑𝑑
𝑃𝑃 = 𝛾𝛾𝛾𝛾 + 𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐 , 𝛾𝛾 = 𝑔𝑔𝑔𝑔 2.25

In the figure a cylindrical element of fluid with cross-


sectional area A and height dz = z2 – z1 is shown. Derive
the equilibrium equation in z-direction. If there is diver
descends from the surface of the sea to a depth of 30 m.
what would be the pressure under which the diver would
be working above that at the surface assuming that the
density of sea water is 1025 Kg m-3 and remains constant?
Force due to P1 on area A acting up = P1A.
Force due to P2 on area A acting down = P2A.
Force due to weight of the element = mg = mass density x g x volume.
= 𝝆𝝆𝝆𝝆𝝆𝝆 𝒅𝒅𝒅𝒅 = 𝝆𝝆𝝆𝝆𝝆𝝆(𝒛𝒛𝟐𝟐 − 𝒛𝒛𝟏𝟏 ).
Since the fluid is at rest, there can be no shear force and, therefore, no vertical
force act on the side of the element due to the surrounding fluid. Taking upward
force as positive and equating the algebraic sum of the force acting to zero.
𝑷𝑷𝟏𝟏 𝑨𝑨 − 𝑷𝑷𝟐𝟐 𝑨𝑨 − 𝝆𝝆𝝆𝝆𝝆𝝆(𝒛𝒛𝟐𝟐 − 𝒛𝒛𝟏𝟏 ) = 𝟎𝟎
𝑷𝑷𝟏𝟏 − 𝑷𝑷𝟐𝟐 = −𝝆𝝆𝝆𝝆(𝒛𝒛𝟐𝟐 − 𝒛𝒛𝟏𝟏 ) (𝒊𝒊)
Thus, in any fluid under gravitational attraction, pressure decreases with
increase of height z. Let z1 = 0, so z2 = 30 and from (i) we get
Increase of pressure = 𝑷𝑷𝟏𝟏 − 𝑷𝑷𝟐𝟐 = −𝝆𝝆𝝆𝝆(𝒛𝒛𝟐𝟐 − 𝒛𝒛𝟏𝟏 )
= -1025 x 9.81 (-30 -0) = 301.7 x 103 Nm-2

2-2-1 Incompressible Fluid


Since the specific weight is equal to the product of fluid density and
acceleration of gravity changes in γ are caused either by a
change in ρ or g. For most engineering applications the variation in g is
negligible, so our main concern is with the possible variation in the fluid
density. For liquids the
variation in density is
usually negligible, even
over large vertical
distances, so that the
assumption of constant
specific weight when
dealing with liquids is a
good one. For this instance, the equation 2.25 can be directly integrated

2.26

2.27

2.28

where h is the distance, which is the depth of fluid measured


downward from the location of p2, This type of pressure distribution is
commonly called a hydrostatic distribution, and Eq. 2.28 shows that in an
incompressible fluid at rest the pressure varies linearly with depth. The
pressure must increase with depth to “hold up” the fluid above it.

As is demonstrated by Eq. 2.27 or 2.28, the pressure in a


homogeneous, incompressible fluid at rest depends on the depth of the
fluid relative to some reference plane, and it is not influenced by the size
or shape of the tank or container in which the fluid is held. Thus, in the
figure, the pressure is the same at all points along the line AB even though
the container may have the very irregular shape shown in the figure. The
actual value of the pressure along AB depends only on the depth, h, the
surface pressure, p0, and the specific weight, γ, of the liquid in the
container.
Because of a leak in a buried
gasoline storage tank, water has
seeped in to the depth shown in
Fig. If the specific gravity of the
gasoline is SG = 0.68, determine
the pressure at the gasoline-water
interface and at the bottom of the
tank. Express the pressure in units
of lb/ft2, lb/in.2, and as a pressure
head in feet of water.
The required equality of pressures at equal elevations throughout a
system is important for the operation of hydraulic jacks, lifts, and presses,
as well as hydraulic controls on aircraft and other types of heavy
machinery. The fundamental idea behind such devices and systems is
demonstrated in Fig. A piston located at one end of a closed system filled
with a liquid, such as oil, can be used to change the pressure throughout
the system, and thus transmit an applied force F1 to a second piston where
the resulting force is F2. Since the pressure p acting on the faces of both
pistons is the same (the effect of elevation changes is usually negligible
for this type of hydraulic device), it follows that
The piston area A2 can be made much larger than A1 and therefore a large
mechanical advantage
can be developed; that
is, a small force applied
at the smaller piston
can be used to develop
a large force at the
larger piston. The applied force could be created manually through some
type of mechanical device, such as a hydraulic jack, or through
compressed air acting directly on the surface of the liquid, as is done in
hydraulic lifts commonly found in service stations.

2-2-2 Compressible Fluid


We normally think of gases such as air, oxygen, and nitrogen as being
compressible fluids since the density of the gas can change significantly
with changes in pressure and temperature. Thus, although Eq. 2.15
applies at a point in a gas, it is necessary to consider the possible
variation in γ before the equation can be integrated. However the specific
weights of common gases are small when compared with those of liquids.
For example, the specific weight of air at sea level and is
whereas the specific weight of water under the same
conditions is Since the specific weights of gases are
comparatively small, it follows from Eq. 2.15 that the pressure gradient in
the vertical direction is correspondingly small, and even over distances of
several hundred feet the pressure will remain essentially constant for a
gas. This means we can neglect the effect of elevation changes on the
pressure in gases in tanks, pipes, and so forth in which the distances
involved are small.
The equation of state for an ideal (or perfect) gas is

2.29

where p is the absolute pressure, R is the gas constant, and T is the


absolute temperature. This relationship can be combined with Eq. 2.15 to
give
2.30

and by separating variables

2.31

where g and R are assumed to be constant over the elevation change from
Although the acceleration of gravity, g, does vary with
elevation, the variation is very small and g is usually assumed constant at
some average value for the range of elevation involved.
Before completing the integration, one must specify the nature of the
variation of temperature with elevation. For example, if we assume that
the temperature has a constant value To over the range
(isothermal conditions), it then follows from Eq. 2.31 that

2.32
This equation provides the desired pressure-elevation relationship for an
isothermal layer. For non-isothermal conditions a similar procedure can
be followed if the temperature-elevation relationship is known, as is
discussed in the following section.
The Empire State Building in New York City, one of the tallest
buildings in the world, rises to a height of approximately 1250 ft.
Estimate the ratio of the pressure at the top of the building to the
pressure at its base, assuming the air to be at a common temperature of
59 0F. Compare this result with that obtained by assuming the air to be
incompressible with γ = 0.0765 lb/ft3 at 14.7 psi (abs) (values for air at
standard conditions).
2-3 Standard Atmosphere
An important application of Eq. 2.31 relates to the variation in pressure
in the earth’s atmosphere. Ideally, we would like to have measurements
of pressure versus altitude over the specific range for the specific
conditions (temperature, reference pressure) for which the pressure is to
be determined. However, this type of information is usually not available.
Thus, a “standard atmosphere” has been determined that can be used in
the design of aircraft, missiles, and spacecraft, and in comparing their
performance under standard conditions. The concept of a standard
atmosphere was first developed in the 1920s, and since that time many
national and international committees and organizations have pursued the
development of such a standard. The currently accepted standard
atmosphere is based on a report published in 1962 and updated in 1976
defining the so-called U.S. standard atmosphere, which is an idealized
representation of middle-latitude, year-round mean conditions of the
earth’s atmosphere. Several important properties for standard atmospheric
conditions at sea level are listed in Table 2.1 shows the temperature
profile for the U.S. standard atmosphere.

Table 2.1 Secondary Properties of U.S. Standard Atmosphere at Sea Level


2-3-1 Measurement of Pressure
Since pressure is a very important characteristic of a fluid field, it is
not surprising that numerous devices and techniques are used in its
measurement. The pressure at a point within a fluid mass will be
designated as either an absolute pressure or a gage pressure. Absolute
pressure is measured
relative to a perfect
vacuum (absolute zero
pressure), whereas
gage pressure is
measured relative to
the local atmospheric
pressure. Thus, a gage
pressure of zero corresponds to a pressure that is equal to the local
atmospheric pressure. Absolute pressures are always positive, but gage
pressures can be either positive or negative depending on whether the
pressure is above atmospheric pressure (a positive value) or below
atmospheric pressure (a negative value). A negative gage pressure is also
referred to as a suction or vacuum pressure.
The measurement of atmospheric pressure is
usually accomplished with a mercury barometer,
which in its simplest form consists of a glass tube
closed at one end with the open end immersed in
a container of mercury as shown in the fig.ure.
The tube is initially filled with mercury (inverted
with its open end up) and then turned upside
down (open end down) with the open end in the
container of mercury. The column of mercury will come to an
equilibrium position where its weight plus the force due to the vapor
pressure (which develops in the space above the column) balances the
force due to the atmospheric pressure. Thus,

2.32

A mountain lake has an average temperature of 10 oC and a maximum


depth of 40 m. For a barometric pressure of 598 mm Hg, determine the
absolute pressure (in Pascal) at the deepest part of the lake.
2-4 Manometry
A standard technique for measuring pressure involves the use of liquid
columns in vertical or inclined tubes. Pressure measuring devices based
on this technique are called manometers. The mercury barometer is an
example of one type of manometer, but there are many other
configurations possible, depending on the particular application. Three
common types of manometers include the piezometer tube, the U-tube
manometer, and the inclined-tube manometer.

2-4-1 Piezometer Tube


The simplest type of manometer consists of a
vertical tube, open at the top, and attached to the
container in which the pressure is desired, as
illustrated in the figure. Since manometers involve
columns of fluids at rest, the fundamental equation
describing their use is Eq.

which gives the pressure at any elevation within a homogeneous fluid in


terms of a reference pressure po and the vertical distance h between p and
po . Remember that in a fluid at rest pressure will increase as we move
downward and will decrease as we move upward. Application of this
equation to the piezometer tube of the above figure indicates that the
pressure pA can be determined by a measurement of h1 through the
relationship where γ1 is the specific weight of the liquid in the container.
Although the piezometer tube is a very simple and accurate pressure
measuring device, it has several disadvantages. It is only suitable if the
pressure in the container is greater than atmospheric pressure (otherwise
air would be sucked into the system), and the pressure to be measured
must be relatively small so the required height of the column is
reasonable. Also, the fluid in the container in which the pressure is to be
measured must be a liquid rather than a gas.

2-4-2 U-Tube Manometer


To overcome the difficulties
noted previously, another type of
manometer which is widely used
consists of a tube formed into the
shape of a U as is shown in the
figure. The fluid in the manometer is
called the gage fluid. To find the
pressure pA in terms of the various
column heights, we start at one end
of the system and work our way around to the other end. Thus, for the U-
tube manometer shown in the figure, we will start at point A and work
around to the open end. The pressure at points A and (1) are the same, and
as we move from point (1) to (2) the pressure will increase by γ1h1. The
pressure at point (2) is equal to the pressure at point (3), since the
pressures at equal elevations in a continuous mass of fluid at rest must be
the same. Note that we could not simply “jump across” from point (1) to
a point at the same elevation in the right-hand tube since these would not
be points within the same continuous mass of fluid. With the pressure at
point (3) specified we now move to the open end where the pressure is
zero. As we move vertically upward the pressure decreases by an amount
γ2h2. In equation form these various steps can be expressed as

Вам также может понравиться